In physics, charge conjugation is a transformation that switches all particles with their corresponding antiparticles, thus changing the sign of all charges: not only electric charge but also the charges relevant to other forces. The term C-symmetry is an abbreviation of the phrase "charge conjugation symmetry", and is used in discussions of the symmetry of physical laws under charge-conjugation. Other important discrete symmetries are P-symmetry (parity) and T-symmetry (time reversal).
These discrete symmetries, C, P and T, are symmetries of the equations that describe the known fundamental forces of nature: electromagnetism, gravity, the strong and the weak interactions. Verifying whether some given mathematical equation correctly models nature requires giving physical interpretation not only to continuous symmetries, such as motion in time, but also to its discrete symmetries, and then determining whether nature adheres to these symmetries. Unlike the continuous symmetries, the interpretation of the discrete symmetries is a bit more intellectually demanding and confusing. An early surprise appeared in the 1950s, when Chien Shiung Wu demonstrated that the weak interaction violated P-symmetry. For several decades, it appeared that the combined symmetry CP was preserved, until CP-violating interactions were discovered. Both discoveries lead to Nobel prizes.
The C-symmetry is particularly troublesome, physically, as the universe is primarily filled with matter, not anti-matter, whereas the naive C-symmetry of the physical laws suggests that there should be equal amounts of both. It is currently believed that CP-violation during the early universe can account for the "excess" matter, although the debate is not settled. Earlier textbooks on cosmology, predating the 1970s, routinely suggested that perhaps distant galaxies were made entirely of anti-matter, thus maintaining a net balance of zero in the universe.
This article focuses on exposing and articulating the C-symmetry of various important equations and theoretical systems, including the Dirac equation and the structure of quantum field theory. The various fundamental particles can be classified according to behavior under charge conjugation; this is described in the article on C-parity.
Charge conjugation occurs as a symmetry in three different but closely related settings: a symmetry of the (classical, non-quantized) solutions of several notable differential equations, including the Klein–Gordon equation and the Dirac equation, a symmetry of the corresponding quantum fields, and in a general setting, a symmetry in (pseudo-)Riemannian geometry. In all three cases, the symmetry is ultimately revealed to be a symmetry under complex conjugation, although exactly what is being conjugated where can be at times obfuscated, depending on notation, coordinate choices and other factors.
A\mu
ei\phi(x)
x.
z=(x+iy)\mapsto\overlinez=(x-iy)
In quantum field theory, charge conjugation can be understood as the exchange of particles with anti-particles. To understand this statement, one must have a minimal understanding of what quantum field theory is. In (vastly) simplified terms, it is a technique for performing calculations to obtain solutions for a system of coupled differential equations via perturbation theory. A key ingredient to this process is the quantum field, one for each of the (free, uncoupled) differential equations in the system. A quantum field is conventionally written as
\psi(x)=\intd3p\sum\sigma,ne-ip ⋅ a\left(\vecp,\sigma,n\right)u\left(\vecp,\sigma,n\right)+eip ⋅ a\dagger\left(\vecp,\sigma,n\right)v\left(\vecp,\sigma,n\right)
where
\vecp
\sigma
n
a
a\dagger
u,v
The creation and annihilation operators obey the canonical commutation relations, in that the one operator "undoes" what the other "creates". This implies that any given solution
u\left(\vecp,\sigma,n\right)
v\left(\vecp,\sigma,n\right)
u\left(\vecp\right)
v\left(\vecp\right)
u
v,
When the fiber to be integrated over is the U(1) fiber of electromagnetism, the dual pairing is such that the direction (orientation) on the fiber is reversed. When the fiber to be integrated over is the SU(3) fiber of the color charge, the dual pairing again reverses orientation. This "just works" for SU(3) because it has two dual fundamental representations
3
\overline3
The above then is a sketch of the general idea of a quantum field in quantum field theory. The physical interpretation is that solutions
u\left(\vecp,\sigma,n\right)
v\left(\vecp,\sigma,n\right)
For general Riemannian and pseudo-Riemannian manifolds, one has a tangent bundle, a cotangent bundle and a metric that ties the two together. There are several interesting things one can do, when presented with this situation. One is that the smooth structure allows differential equations to be posed on the manifold; the tangent and cotangent spaces provide enough structure to perform calculus on manifolds. Of key interest is the Laplacian, and, with a constant term, what amounts to the Klein–Gordon operator. Cotangent bundles, by their basic construction, are always symplectic manifolds. Symplectic manifolds have canonical coordinates
x,p
A second interesting thing one can do is to construct a spin structure. Perhaps the most remarkable thing about this is that it is a very recognizable generalization to a
(p,q)
\gamma5
SO(p,q)
SO(1,3)
SpinC(p,q).
SO(p,q) x U(1).
The
U(1)
F=dA
A
U(1)
U(1)
What remains is to work through the discrete symmetries of the above construction. There are several that appear to generalize P-symmetry and T-symmetry. Identifying the
p
q
p
q
There are two ways to react to this. One is to treat it as an interesting curiosity. The other is to realize that, in low dimensions (in low-dimensional spacetime) there are many "accidental" isomorphisms between various Lie groups and other assorted structures. Being able to examine them in a general setting disentangles these relationships, exposing more clearly "where things come from".
The laws of electromagnetism (both classical and quantum) are invariant under the exchange of electrical charges with their negatives. For the case of electrons and quarks, both of which are fundamental particle fermion fields, the single-particle field excitations are described by the Dirac equation
(i{\partial /}-q{A /}-m)\psi=0
One wishes to find a charge-conjugate solution
(i{\partial /}+q{A /}-m)\psic=0
A handful of algebraic manipulations are sufficient to obtain the second from the first. Standard expositions of the Dirac equation demonstrate a conjugate field
\overline\psi=\psi\dagger\gamma0,
\overline\psi(-i{\partial /}-q{A /}-m)=0
Note that some but not all of the signs have flipped. Transposing this back again gives almost the desired form, provided that one can find a 4×4 matrix
C
C-1\gamma\muC=
sf{T} | |
-\gamma | |
\mu |
The charge conjugate solution is then given by the involution
\psi\mapsto
c=η | |
\psi | |
c |
C\overline\psisf{T}
The 4×4 matrix
C,
ηc
|ηc|=1,
ηc=1.
The interplay between chirality and charge conjugation is a bit subtle, and requires articulation. It is often said that charge conjugation does not alter the chirality of particles. This is not the case for fields, the difference arising in the "hole theory" interpretation of particles, where an anti-particle is interpreted as the absence of a particle. This is articulated below.
Conventionally,
\gamma5
C\gamma5C-1=
sf{T} | |
\gamma | |
5 |
and whether or not
sf{T} | |
\gamma | |
5 |
\gamma5
sf{T}= | |
\gamma | |
5 |
\gamma5
sf{T}= | |
\gamma | |
5 |
-\gamma5
For the case of massless Dirac spinor fields, chirality is equal to helicity for the positive energy solutions (and minus the helicity for negative energy solutions). One obtains this by writing the massless Dirac equation as
i\partial /\psi=0
Multiplying by
\gamma5\gamma0=-i\gamma1\gamma2\gamma3
{\epsilonij
where
\sigma\mu\nu=i\left[\gamma\mu,\gamma\nu\right]/2
\epsilonijk
\Sigmam\equiv{\epsilonij
\psi(x)=e-ik ⋅ \psi(k)
k ⋅ k=0
{\hatk}i=ki/k0
\left(\Sigma ⋅ \hatk\right)\psi=\gamma5\psi~.
Examining the above, one concludes that angular momentum eigenstates (helicity eigenstates) correspond to eigenstates of the chiral operator. This allows the massless Dirac field to be cleanly split into a pair of Weyl spinors
\psiL
\psiR,
\left(-p0+\sigma ⋅ \vecp\right)\psiR=0
and
\left(p0+\sigma ⋅ \vecp\right)\psiL=0
Note the freedom one has to equate negative helicity with negative energy, and thus the anti-particle with the particle of opposite helicity. To be clear, the
\sigma
p\mu=i\partial\mu
Taking the Weyl representation of the gamma matrices, one may write a (now taken to be massive) Dirac spinor as
\psi=\begin{pmatrix}\psiL\ \psiR\end{pmatrix}
The corresponding dual (anti-particle) field is
\overline{\psi}sf{T} =\left(\psi\dagger\gamma0\right)sf{T} =\begin{pmatrix}0&I\ I&0\end{pmatrix}\begin{pmatrix}
* | |
\psi | |
L |
* | |
\ \psi | |
R |
\end{pmatrix} =\begin{pmatrix}
* | |
\psi | |
R |
* | |
\ \psi | |
L |
\end{pmatrix}
The charge-conjugate spinors are
\psic =\begin{pmatrix}
c | |
\psi | |
R |
\end{pmatrix}=ηcC\overline\psisf{T} =ηc\begin{pmatrix}-i\sigma2&0\ 0&i\sigma2\end{pmatrix}\begin{pmatrix}
* | |
\psi | |
R |
* | |
\ \psi | |
L |
\end{pmatrix} =ηc\begin{pmatrix}
* | |
-i\sigma | |
R |
* | |
\ i\sigma | |
L |
\end{pmatrix}
where, as before,
ηc
ηc=1.
\psi\left(t,\vecx\right)\mapsto\psip\left(t,\vecx\right)=\gamma0\psi\left(t,-\vecx\right)
Under combined charge and parity, one then has
\psi\left(t,\vecx\right)\mapsto\psicp\left(t,\vecx\right)=\begin{pmatrix}
cp | |
\psi | |
L |
\left(t,\vec
cp | |
x\right)\ \psi | |
R |
\left(t,\vecx\right)\end{pmatrix}=ηc\begin{pmatrix}
*\left(t, | |
-i\sigma | |
L |
-\vecx\right)
*\left(t, | |
\ i\sigma | |
R |
-\vecx\right)\end{pmatrix}
Conventionally, one takes
ηc=1
The Majorana condition imposes a constraint between the field and its charge conjugate, namely that they must be equal:
\psi=\psic.
Doing so requires some notational care. In many texts discussing charge conjugation, the involution
\psi\mapsto\psic
l{C}
C:\psi\mapsto\psic
C\psi=\psic.
C\psi=\psi.
C\psi(\pm)=\pm\psi(\pm).
\psi(+)=\begin{pmatrix}
* | |
\psi | |
L |
\end{pmatrix}
and
\psi(-)=\begin{pmatrix}
*\ \psi | |
i\sigma | |
R |
\end{pmatrix}
The Majorana spinor is conventionally taken as just the positive eigenstate, namely
\psi(+).
\gamma5
\gamma5C=-C\gamma5
This is readily verified by direct substitution. Bear in mind that
C
The projectors onto the chiral eigenstates can be written as
PL=\left(1-\gamma5\right)/2
PR=\left(1+\gamma5\right)/2,
PLC=CPR~.
This directly demonstrates that charge conjugation, applied to single-particle complex-number-valued solutions of the Dirac equation flips the chirality of the solution. The projectors onto the charge conjugation eigenspaces are
P(+)=(1+C)PL
P(-)=(1-C)PR.
The phase factor
ηc
ηc
l{C}
\psi\mapsto\psic=l{C} \psi l{C}\dagger=
sf{T} | |
η | |
c C \overline\psi |
\overline\psi\mapsto\overline\psic=l{C} \overline\psi l{C}\dagger=
sf{T} C | |
η | |
c \psi |
-1
A\mu\mapsto
c | |
A | |
\mu |
=
\dagger | |
l{C} A | |
\mu l{C} |
=-A\mu
where the non-calligraphic
C
Charge conjugation does not alter the chirality of particles. A left-handed neutrino would be taken by charge conjugation into a left-handed antineutrino, which does not interact in the Standard Model. This property is what is meant by the "maximal violation" of C-symmetry in the weak interaction.
Some postulated extensions of the Standard Model, like left-right models, restore this C-symmetry.
The Dirac field has a "hidden"
U(1)
C
U(1)
One very conventional technique is simply to start with two real scalar fields,
\phi
\chi
\psil\stackrel{def
The charge conjugation involution is then the mapping
C:i\mapsto-i
C:\phi\mapsto\phi
C:\chi\mapsto\chi
C:\psi\mapsto\psi*.
\mapsto
C\phi=\phi
C\chi=\chi
C\psi=\psi*.
The above describes the conventional construction of a charged scalar field. It is also possible to introduce additional algebraic structure into the fields in other ways. In particular, one may define a "real" field behaving as
C:\phi\mapsto-\phi
C:\psi\mapsto-\psi*.
In physics literature, a transformation such as
C:\phi\mapsto\phic=-\phi
\phi
R x Z2
Z2=\{+1,-1\}.
\phi=(r,c)
C:(r,c)\mapsto(r,-c).
It was believed for some time that C-symmetry could be combined with the parity-inversion transformation (see P-symmetry) to preserve a combined CP-symmetry. However, violations of this symmetry have been identified in the weak interactions (particularly in the kaons and B mesons). In the Standard Model, this CP violation is due to a single phase in the CKM matrix. If CP is combined with time reversal (T-symmetry), the resulting CPT-symmetry can be shown using only the Wightman axioms to be universally obeyed.
The analog of charge conjugation can be defined for higher-dimensional gamma matrices, with an explicit construction for Weyl spinors given in the article on Weyl–Brauer matrices. Note, however, spinors as defined abstractly in the representation theory of Clifford algebras are not fields; rather, they should be thought of as existing on a zero-dimensional spacetime.
The analog of T-symmetry follows from
\gamma1\gamma3
This is done by taking the tangent space as a vector space, extending it to a tensor algebra, and then using an inner product on the vector space to define a Clifford algebra. Treating each such algebra as a fiber, one obtains a fiber bundle called the Clifford bundle. Under a change of basis of the tangent space, elements of the Clifford algebra transform according to the spin group. Building a principle fiber bundle with the spin group as the fiber results in a spin structure.
All that is missing in the above paragraphs are the spinors themselves. These require the "complexification" of the tangent manifold: tensoring it with the complex plane. Once this is done, the Weyl spinors can be constructed. These have the form
wj=
1 | |
\sqrt{2 |
where the
ej
V=TpM
p\inM
M.
V ⊗ C=W ⊕ \overlineW
The alternating algebra
\wedgeW